Spectroscopic and Kinetic Evidence for an Accumulating Intermediate

Apr 4, 2007 - ... rates of product formation and disappearance of the substrate were the same, i.e., ... Miriam Karni, Claude F. Bernasconi, and Zvi R...
0 downloads 0 Views 184KB Size
Spectroscopic and Kinetic Evidence for an Accumulating Intermediate in an SNV Reaction with Amine Nucleophiles. Reaction of Methyl β-Methylthio-r-nitrocinnamate with Piperidine and Morpholine Claude F. Bernasconi,*,† Shoshana D. Brown,† Irina Eventova,‡ and Zvi Rappoport‡ Department of Chemistry and Biochemistry, UniVersity of California, Santa Cruz, California 95064, and Department of Organic Chemistry, The Hebrew UniVersity of Jerusalem, Jerusalem 91904, Israel [email protected] ReceiVed December 18, 2006

A spectroscopic and kinetic study of the reaction of methyl β-methylthio-R-nitrocinnamate (4-SMe) with morpholine, piperidine, and hydroxide ion in 50% DMSO/50% water (v/v) at 20 °C is reported. The reactions of 4-SMe with piperidine in a pH range from 10.12 to 11.66 and those with morpholine at pH 12.0 are characterized by two kinetic processes when monitored at λmax (364 nm) of the substrate, but by only one process when monitored at λmax (388) nm of the product. The rate constants obtained at 388 nm were the same as those determined for the slower of the two processes at 364 nm. These rate constants refer to product formation, whereas the faster process observed at 364 nm is associated with the loss of reactant to form an intermediate. In contrast, for the reaction of 4-SMe with morpholine at pH 8.62 the rates of product formation and disappearance of the substrate were the same, i.e., there is no accumulation of an intermediate. Likewise, the reaction of 4-SMe with OH- did not yield a detectable intermediate. The factors that allow the accumulation of intermediates in certain SNV reactions but not in others are discussed in detail, and structure-reactivity comparisons are made with reactions of piperidine and morpholine with other highly activated vinylic substrates.

Introduction The nucleophilic vinylic substitution (SNV) on substrates activated by electron-withdrawing groups is a stepwise process as shown in eq 1 for the reaction with an anionic nucleophile; X and Y are the activating substituents and LG- is the leaving group.1 Early kinetic studies provided mainly indirect evidence for the stepwise nature of the reaction, but more recently several † ‡

University of California. The Hebrew University.

(1) For reviews, see: (a) Rappoport, Z. AdV. Phys. Org. Chem. 1969, 7, 1. (b) Modena, G. Acc. Chem. Res. 1971, 4, 73. (c) Miller, S. I. Tetrahedron 1977, 33, 1211. (d) Rappoport, Z. Acc. Chem. Res. 1981, 14, 7. (e) Rappoport, Z. Recl. TraV. Chim. Pays-Bas 1985, 104, 309. (f) Shainyan, B. A. Usp. Khim. 1986, 55, 942. (g) Rappoport, Z. Acc. Chem. Res. 1992, 25, 474.

systems were found where the intermediate actually accumulates to detectable levels, allowing an experimental determination of the individual rate constants k1, k-1, and k2.2 Some notable examples are the reactions of 1-OMe,3,4 1-SMe,4 2-OMe,5 (2) Bernasconi, C. F.; Ketner, R. J.; Ragains, M. L.; Chen, X.; Rappoport, Z. J. Am. Chem. Soc. 2001, 123, 2155. 10.1021/jo062602s CCC: $37.00 © 2007 American Chemical Society

3302

J. Org. Chem. 2007, 72, 3302-3310

Published on Web 04/04/2007

SNV Reaction with Amine Nucleophiles

2-SMe,5 3-OMe,2 and 4-SMe2 with thiolate ions as well as some other substrates.2

There are two requirements for an intermediate such as TNu to become detectable. The first is that the equilibrium of the first step is favorable, i.e., (k1/k-1)[Nu-] ) K1[Nu-] > (.) 1 (“thermodynamic condition”). The second is that the formation of TNu is faster than its conversion to products, i.e., k1[Nu ] > (.) k2 (“kinetic condition”). The first condition can be met by a combination of a strong nucleophile and strongly electron-withdrawing substituents, whereas for the second condition there is the additional requirement that LG- is a sluggish leaving group. All of the above-mentioned examples meet both requirements. For the reactions with amine nucleophiles the mechanism involves additional steps due to the acidic nature of the initially formed zwitterionic intermediate as shown in eq 2; the

only cases involving such amines studied so far, have yielded detectable intermediates, although paradoxically the reaction of 1-OMe with the much more weakly basic methoxyamine (pKa ) 4.70) and N-methylmethoxyamine (pKa ) 4.67) did allow 9 the direct observation of TA. The reason why the intermediates in the reactions of 1-OMe and 2-SMe with strongly basic amines were undetectable is not that the thermodynamic condition was not met-it was amply met-it is the kinetic condition that remained elusive. This is because leaving group departure is strongly accelerated by the 9 electronic push from the nitrogen lone pair in TA , leading to transition state stabilization that results from the developing product resonance (eq 5).

The reason why TA was observed in the reaction of 1-OMe with weakly basic amines is that the push shows a much stronger dependence on amine basicity (βpush ) 0.71)9 than the nucleophilic addition step (βnuc ) 0.25).9 This means that the reduced nucleophilic reactivity resulting from the lower basicity is more than offset by the decreased acceleration of the product forming step. In this paper we report spectroscopic and kinetic evidence that in the reaction of methyl β-methylthio-R-nitrocinnamate (4-SMe) with piperidine and morpholine TA accumulates to detectable levels under certain conditions. This is the first observed example of a nucleophilic vinylic substitution by strongly basic amines that leads to such intermediate accumulation. We also report kinetic data on the hydrolysis of 4-SMe in basic solution. Results

+ kAH 3 [R′R′′NH2 ] term refers to general acid (AH)-catalyzed leaving group departure by the protonated amine, while kH3 2O refers to noncatalyzed or water-catalyzed loss of the leaving group. In this case the conditions for detectability of either one or both intermediates (T( A , TA ) are given by eq 3 (thermodynamic condition) and eq 4 (kinetic condition), respectively:6

(K1 + K1K( a /aH+)[R′R′′NH] > (.) 1

(3)

( H2O + k1[R′R′′NH] > (.) {K( a /(Ka + aH+)}(k3 + kAH 3 [R′R′′NH2 ]) (4)

Interestingly, the reactions of neither 1-OMe7 nor 2-SMe8 with strongly basic amines such as piperidine or n-butylamine, the (3) Bernasconi, C. F.: Fassberg, J.; Killion, R. B., Jr.; Rappoport, Z. J. Am. Chem. Soc. 1989, 111, 6862. (4) Bernasconi, C. F.; Fassberg, J.; Killion, R. B., Jr.; Rappoport, Z. J. Am. Chem. Soc. 1990, 112, 3169.

General Features. The synthesis of 4-SMe led to a mixture of E and Z isomers with an E/Z ratio of 0.5. Attempts at separating the mixture were unsuccessful. Our results suggest that the reactivities of the two isomers are very similar and do not lead to complications such as biphasic kinetics. The reaction of 4-SMe with piperidine and morpholine leads to 4-Pip and 4-Mor, respectively. Both 4-Pip and 4-Mor have been thoroughly characterized, including 1H NMR, 13C NMR, MS, an X-ray crystallographic structure of 4-Pip, and UV data.10 (5) (a) Bernasconi, C. F.; Ketner, R. J.; Chen, X.; Rappoport, Z. J. Am. Chem. Soc. 1998, 120, 7461. (b) Bernasconi, C. F.; Ketner, R. B.; Brown, S. D.; Chen, X.; Rappoport, Z. J. Org. Chem. 1999, 64, 8829. (c) Bernasconi, C. F.; Ketner, R. J.; Chen, X.; Rappoport, Z. Can. J. Chem. 1999, 77, 584. (6) Usually a pH higher than pK( a is required; under such conditions eqs + 3 and 4 simplify to (K1K( a /aH )[R′R′′NH] > (.) 1 and k1[R′R′′NH] > + (.) kH3 2O + kAH 3 [R′R′′NH2 ]. (7) Bernasconi, C. F.; Fassberg, J.; Killion, R. B., Jr.; Rappoport, Z. J. Org. Chem. 1990, 55, 4568. (8) (a) Bernasconi, C. F.; Ali, M.; Nguyen, K.; Ruddat, V.; Rappoport, Z. J. Org. Chem. 2004, 69, 9249. (b) Ali, M.; Biswas, S.; Rappoport, Z.; Bernasconi, C. F. J. Phys. Org. Chem. 2006, 19, 647. (9) (a) Bernasconi, C. F.; Leyes, A. E.; Rappoport, Z.; Eventova, I. J. Am. Chem. Soc. 1993, 115, 7513. (b) Bernasconi, C. F.; Leyes, A. E.; Eventova, I.; Rappoport, Z. J. Am. Chem. Soc. 1995, 117, 1703. (10) Beit-Yannai, M.; Chen, X.; Rappoport, Z. J. Chem. Soc., Perkin Trans. 2 2001, 1534.

J. Org. Chem, Vol. 72, No. 9, 2007 3303

Bernasconi et al.

FIGURE 1. Time-resolved spectra of the reaction of 4-SMe with morpholine at pH 8.62 and [morpholine] ) 0.20 M; time intervals 20 s.

FIGURE 2. Time-resolved spectra of the reaction of 4-SMe with piperidine at pH 11.0 and [piperidine] ) 0.19 M; time intervals 20 s.

The latter show λmax in acetonitrile at 274/388 nm for 4-Pip and at 275/386 nm for 4-Mor, respectively.

All reactions were conducted in 50% DMSO/50% water (v/v) at 20 °C and an ionic strength of 0.5 M maintained with KCl. The kinetic experiments were conducted under pseudofirst-order conditions with 4-SMe as the minor component. Reaction of 4-SMe with Morpholine and Piperidine. UV/ vis Spectra. When the reaction with morpholine was conducted at a pH close to the pKa of the morpholinium ion (8.72), there was a clean transformation of reactants to products. This is apparent from Figure 1, which shows two sharp isosbestic points in the time-resolved UV/vis spectra taken at pH 8.62. It implies that none of the two assumed intermediates (T( A and TA ) accumulates to detectable levels, i.e., they behave as steady state intermediates. As shown in Figure 2, the situation is different for the reaction with piperidine at a pH close to the pKa of the piperidinium ion 3304 J. Org. Chem., Vol. 72, No. 9, 2007

FIGURE 3. Reaction of 4-SMe with morpholine at pH 8.76: (O) 388 k364 obsd; (b) kobsd.

(11.02). In this case the absence of sharp isosbestic points indicates the presence of at least one non-steady-state intermediate. A similar behavior was observed when the reaction with morpholine was run in a triethylamine buffer at pH 11.96 or in a KOH solution at pH 12.0, again suggesting the presence of a non-steady-state intermediate (spectra not shown). Kinetics. The spectra shown in Figure 1 for the reaction with morpholine suggest that the rate of product formation measured at 388 nm should be the same as the rate of disappearance of the reactant determined at 364 nm. Plots of the observed pseudo364 first-order rate constants, k388 obsd and kobsd, versus morpholine concentration at pH 8.76 shown in Figure 3 confirm this expectation.11 The situation is different at pH 12.0; here product formation is slower than loss of reactant, as expected on the basis of the spectral observations. Kinetic measurements at 388 nm provided rate constants for product formation, k388 obsd. Kinetic measurements at 364 nm yielded two rate constants: k364 obsd(R) for the fast loss of reactant associated with a decrease in absorption and k364 obsd(P) for the slower rate of product formation associated with a small increase in absorption. Because the absorbance changes associated with k364 obsd(P) were , the data obtained at smaller than the ones associated with k388 obsd 388 nm are deemed somewhat more reliable. Figure 4 shows 364 plots of k388 obsd and kobsd(R) versus morpholine concentration. Slower rates of product formation compared to the rate of substrate loss were also observed for the reaction with piperidine at several pH values. Figures S1 and S2 in Supporting Information12 show representative plots of k364 obsd(R) versus piperidine concentration, and Figures S3 and S412 show corresponding plots of k388 obsd; for reasons explained in Discussion, the latter data are plotted versus piperidinium ion rather than piperidine concentration. As with the morpholine reaction, rates of product formation could also be determined at 364 nm, but the k388 obsd values are again deemed more reliable than the (P) values because of smaller absorbance changes and also k364 obsd 364 because the separation between k364 obsd(R) and kobsd(P) was not very large. 364 (11) The k388 obsd values are slightly larger than the kobsd values, which we attribute to experimental error; if there was a slight accumulation of an 364 intermediate one would expect k388 obsd to be smaller than kobsd. (12) See Supporting Information paragraph at the end of this paper.

SNV Reaction with Amine Nucleophiles

FIGURE 4. Reaction of 4-SMe with morpholine at pH 12.0: (O) 388 k364 obsd(R); (b) kobsd.

FIGURE 5. Reaction of 4-SMe with piperidine. Plots according to eq 9: (O) 364 nm; (b) 388 nm.

Hydrolysis of 4-SMe. Rates of hydrolysis were measured in KOH solutions ranging from 0.01 to 0.35 M. The reaction was monitored by following the disappearance of the substrate at 306 and 364 nm. The pseudo-first-order rate constants are linearly dependent on [KOH] (Figure S5)12 and are consistent with eq 6:

kobsd ) kH2O + kOH[OH-]

(6)

The slope yielded kOH ) 0.415 ( 0.011 M-1 s-1 from the data obtained at 364 nm and kOH ) 0.343 ( 0.011 M-1 s-1 at 306 nm. The former is deemed more reliable because the absorbance changes at 364 nm were much larger than those at 306 nm. The intercepts of these plots were too small to yield a reliable value for kH2O. The hydrolysis reaction is followed by two kinetic processes on much slower time scales; they lead to unidentified products and were not further investigated. Discussion Interpretation of Kinetic Data. Under the conditions where one or both intermediates are detectable, the fast process refers to the reversible formation of T( A and TA , while the slower process refers to the conversion of TA (in fast equilibrium with 364 T( A ) to products. Hence kobsd(R) is approximated by eq 7 and 388 364 kobsd ) kobsd(P) are approximated by eq 8:

k364 obsd(R)

) k1[R′R′′NH] + k-1

364 k388 obsd ) kobsd(P) )

a H+ K( a + aH+

(7)

K( a

+ (kH3 2O + kAH 3 [R′R′′NH2 ]) K( + a + a H

(8)

We use the term “approximated” because the time separation between the fast and slow process was, in most cases, significantly less than a factor of 10 and as low as a factor of 3 or 4, especially for the reaction with piperidine at pH 10.12 and 10.30. This leads to some kinetic coupling between the two processes and reduces the accuracy of the kobsd values. However,

for our purposes eqs 7 and 8 are adequate. The situation described by eqs 7 and 8 pertains to the reaction of 4-SMe with piperidine under all conditions examined in this study and to the reaction with morpholine at pH 12.0 in KOH solutions and at pH 11.96 in dilute triethylamine buffers. The plots of k364 obsd(R) versus amine concentration do not yield well-defined intercepts, suggesting that the k-1aH+/(K( a + aH+) term is negligible. The slopes of these plots provide k1 values; the fact that for the piperidine reaction these slopes are essentially pH-independent (Table S1)12 supports our interpretation in terms of eq 7. 364 The plots of k388 obsd or kobsd(P) also have negligible intercepts, indicating that the water catalyzed leaving group departure (kH3 2O) is negligible compared to the ammonium ion catalyzed conversion of TA to products. The pH dependence of the slopes of these plots examined for the piperidine reaction show a nonlinear increase with increasing pH (Table S2);12 a plot of slope-1 versus aH+ according to eq 9 is shown in Figure 5; it -2 M-1 s-1 and pK( ) 11.08 yields kAH 3 ) (4.05 ( 0.98) × 10 a AH ( 0.12 at 364 nm, and k3 ) (4.88 ( 2.85) × 10-2 M-1 s-1 and pK( a ) 11.30 ( 0.22 at 388 nm.

slope-1 )

1

+ AH

k3

aH+ AH K( a k3

(9)

For the morpholine reaction conducted at pH 12.0 we can safely assume pH . pK( a since in similar situations ( one always finds pK( a (piperidine) - pKa (morpholine) ≈ AH AH pKa (piperidine) - pKa (morpholine).3,4,13 Hence eq 8 simplifies to eq 10. The kAH 3 value determined from the slope -1 s-1. of k364 obsd(P) is 2.08 ( 0.23 M 364 AH + k388 obsd ) kobsd(P) ) k3 [R′R′′NH2 ]

(10)

For the reaction of 4-SMe with morpholine at pH 8.76 there is no intermediate accumulation, and hence the steady state 388 applies and k364 obsd ) kobsd are given by eq 11. (13) Bernasconi, C. F. Tetrahedron 1989, 45, 4017.

J. Org. Chem, Vol. 72, No. 9, 2007 3305

Bernasconi et al. TABLE 1. Summary of Kinetic Parameters for the Reactions of

364 k388 obsd ) kobsd ) ( kH3 2OK( kAH 3 Ka a [R′R′′NH] + [OH-] Kw KAH a [R′R′′NH] k1 ( kAH kH3 2OK( 3 Ka a k-1 + AH [R′R′′NH] + [OH ] Kw Ka

4-SMe with Piperidine, Morpholine, and OH- in 50% DMSO/50% Water (v/v) at 20 °C 364 nm

(11)

388 The plots of k364 obsd and kobsd (Figure 3) show some upward curvature, which is consistent with catalysis by morpholine at ( AH low concentrations and arises from the (kAH 3 Ka /Ka )[R′R′′NH] term in eq 11. As the morpholine concentration increases, the curvature decreases. These results imply that at low amine ( concentration the relationship of eq 12 holds with the (kAH 3 Ka / AH 14 Ka )[R′R′′NH] term being dominant.

388 nm

piperidine (pKAH a ) 11.02) (9.63 ( 1.55) × 10-2a (4.05 ( 0.98) × 10-2 (4.88 ( 2.85) × 10-2

k1, M-1 s-1 -1 s-1 kAH 3 ,M k1/kAH 3 11.08 ( 0.12 pK( a

average

11.30 ( 0.22

9.63 × 10-2 4.46 × 10-2 2.16 11.19

morpholine (pKAH a ) 8.72) k1, M-1 s-1 (1.71 ( 0.17) × 10-2b AH k3 , M-1 s-1 2.08 ( 0.23 k1/kAH 3 pK( a

1.71 × 10-2 2.08 8.22 × 10-3 ≈8.88c

k1, M-1 s-1

0.415 ( 0.011

OH0.415

a

( kH3 2OK( kAH 3 Ka a [R′R′′NH] + [OH- ] j k-1 AH K Ka w

(12)

At the higher concentrations the decrease in the upward curvature means that either eq 13 or eq 14 holds. ( AH (kAH 3 Ka /Ka )[R′R′′NH] J k-1

(13)

( AH (kAH 3 Ka /Ka )[R′R′′NH] . k-1

(14)

We prefer the former interpretation, because if eq 14 were valid, it would imply that eq 11 reduces to eq 15 and that the slope of the quasi linear portion of the plots in Figure 3 are equal to k1. 364 k388 obsd ) kobsd ) k1[R′R′′NH]

(15)

However, these slopes are ∼7.0 × 10-3 M-1 s-1 at 364 nm and ≈8.2 × 10-3 M-1 s-1 at 388 nm, respectively; they are smaller than the k1 value of (1.71 ( 0.17) × 10-2 M-1 s-1 determined at pH 12.0. Hence, the condition of eq 14 has not been reached. Summary of Rate Constants and Comparisons with Other Systems. Table 1 provides a summary of the kinetic parameters ( obtained in this study. Note that the k1, kAH 3 , and pKa values for the piperidine reaction are associated with relatively large experimental uncertainties; as pointed out earlier, this is most likely the result of coupling between the fast and slow processes that leads to inaccuracies in the kobsd values. The parameters summarized in Table 2 allow comparisons to be made among reactions of similar substrates with the same nucleophiles. Table 2 includes the pKCH values of the respective parent carbon a acids (5, 6, and 7, respectively) and log kPT ° values, which refer

(14) Since for the piperidine reaction the kAH 3 pathway is dominant over the kH3 2O pathway, the same must be true for the morpholine reaction, especially since the morpholinium ion should be a stronger acid catalyst than the piperidinium ion.

3306 J. Org. Chem., Vol. 72, No. 9, 2007

Average from slopes at six pH values (Table S1). b Average from slopes ( at two pH values (Table S1). c Estimated as pK( a (mor) ≈ pKa (pip) AH (pip) + pK (mor). pKAH a a

to the intrinsic rate constants15 for the proton transfer from these carbon acids (5, 6, and 7) to secondary alicyclic amines. The pKCH values may be regarded as an approximate measure of a the relative thermodynamic stabilities of the respective intermediates; the log kPT ° values serve as rough indicators of the relative intrinsic rate constants for nucleophilic attack on the respective vinylic substrates, except for possible distortions by other effects as will be discussed below. The following points are noteworthy: (1) In the reactions of 4-SMe with piperidine and morpholine, leaving group departure catalyzed by the respective protonated amine (kAH 3 ) is the dominant pathway while the noncatalyzed or water-catalyzed pathway (kH3 2O) is negligible. The kAH 3 (mor)/ (pip) ratio of 48.9, which corresponds to a Brønsted R) kAH 3 0.73, suggests a transition state where proton transfer is well advanced. The dominance of the kAH 3 pathway is reminiscent of the reaction of 2-SMe8a with piperidine where kAH 3 is 433-fold larger than kH3 2O, but different from the situation in the reaction of 1-OMe with amines where it is the kH3 2O pathway that is dominant.7,9 The fact that the reactions with the MeS- leaving group are more sensitive to acid catalysis than the reaction with the MeO- leaving group is counterintuitive because the methoxide ion is more basic than the methylthiolate ion and usually reactions involving oxyanions as leaving groups are more prone to acid catalysis than those involving thiolate ions.17 A possible explanation for the absence of significant acid catalysis of MeOexpulsion from TA derived from 1-OMe lies in the very strong electronic push (eq 5) that prevails in the noncatalyzed MeOexpulsion (kH3 2O, βnuc ) 0.71). This makes the kH3 2O step competitive with the acid-catalyzed pathway because the transition state of the acid-catalyzed pathway is probably too crowded for the optimal coplanarity required for an effective push to apply. In the reactions of 2-SMe and 4-SMe there is (15) The intrinsic rate constants refer to rate constants that have been interpolated or extrapolated from Brønsted plots to ∆pKa + log(p/q) ) 0 CH where ∆pKa ) pKAH a - pKa is the difference between the pKa values of the protonated amines and the carbon acids, respectively, and p and q refer to statistical factors.16 (16) Keeffe, J. R.; Kresge, A. J. In InVestigation of Rates and Mechanisms of Reactions; Bernasconi, C. F., Ed.; Wiley-Interscience: New York, 1986; Part 1, p 747. (17) Bernasconi, C. F.; Ketner, R. J.; Brown, S. D.; Chen, X.; Rappoport, Z. J. Org. Chem. 1999, 64, 8829.

SNV Reaction with Amine Nucleophiles TABLE 2. Comparison of Kinetic Parameters of the Reactions of 4-SMe with the Reactions of 1-OMe and 2-SMe

a Reference 7. b Reference 31. c Bernasconi, C. F.; Kliner, D. A. V.; Mullin, A. S.; Ni, J. X. J. Org. Chem. 1988, 53, 3342. d Bernasconi, C. F.; Pe ´ rezLorenzo, M.; Brown, S. D. Unpublished results. e Reference 8a. f Estimated, see ref 8a. g Reference 5c. h Bernasconi, C. F.; Oliphant, N. Unpublished results. i pK of parent carbon acid; see text. j kPT is the intrinsic rate constant for proton transfers from parent carbon acid; see text. a o

already too much crowding even at the transition state of the non-catalyzed reaction because of the large size of the MeS group. This leads to a reduction of the push in the kH3 2O step. We shall return to these points when dealing with the question of why TA is observable in the reaction of 2-SMe with piperidine and morpholine but not in the reaction of 1-OMe with the same amines. (2) The pK( a (pip) value for 4-SMe is somewhat higher than that for 2-SMe, presumably because the O2NCCO2Me moiety is somewhat less electron-withdrawing than the C(COO)2C(CH3)2 moiety, as suggested by the respective pKCH a values (Table 2). This contrasts with the much lower pK( a (pip) value for 1-SMe. In order to develop some understanding of this contrast it is useful to look at the pK( a values of the piperidine adducts of olefinic substrates such as 8, 9, and 10 that lack a leaving group. Two factors influence these pK( a

values. One is the electron-withdrawing strength of the activating substituents, which is strongest in the case of 9, weakest in the case of 10 (the pKCH of acetylacetone is 9.1218,23), and a intermediate in the case of 8. The second and more important factor is intramolecular hydrogen bonding between the ammonium ion and the negative oxygens exemplified by 11. This hydrogen bond reduces the acidity of the ammonium ion and (18) In 50% DMSO/50% water (v/v) at 20°C. (19) Bernasconi, C. F.; Renfrow, R. A. J. Org. Chem. 1987, 52, 3035. (20) In water at 25°C. (21) Bernasconi, C. F.; Murray, C. J. J. Am. Chem. Soc. 1986, 108, 5251.

is strongest in the case of 10, intermediate in the case of 9, and weakest for T( A derived from 8. This implies that anionic carbonyl oxygens are better hydrogen acceptors than anionic nitro group oxygens. A possible reason for this is that there is repulsion between the positive charges on the ammonium nitrogen and the nitrogen of the nitro group that (partially) offsets the stabilizing effect of hydrogen bonding. ( Turning to the pK( a values of TA derived from 1-OMe, 2-SMe, and 4-SMe (Table 2), we note the following. For 1-OMe there is a reduction by 2 pK units relative to that for 8, which may be accounted for by the electron-withdrawing inductive effect of the MeO group. In 2-SMe the degree of reduction relative to the pK( a of 9 is unknown but probably less than 2 pK units since the MeS group is less electron-withdrawing than the MeO group.24 The pK( a of 4-SMe (11.19) hence fits in well with the notion that there is substantial intramolecular hydrogen bonding in T( A derived not only from 2-SMe but also from 4-SMe and that this hydrogen bonding is most likely to the carbonyl oxygen of the ester group rather than to an oxygen of the nitro group. (3) In comparing 4-SMe to 2-SMe one would expect that 4-SMe should be less reactive toward nucleophilic attack than suggests a lower electrophi2-SMe because the higher pKCH a licity or lower stability of the respective intermediates and the lower log kPT ° implies a lower rate constant even if the stabilities of the respective intermediates were the same. The (22) Bernasconi, C. F.; Kanavarioti, A. J. Am. Chem. Soc. 1986, 108, 7744. (23) Bernasconi, C. F.; Bunnell, R. D. Isr. J. Chem. 1985, 26, 420. (24) σF(OMe) ) 0.30; σF(SMe) ) 0.20.25

J. Org. Chem, Vol. 72, No. 9, 2007 3307

Bernasconi et al.

fact that k1(OH-) for 4-SMe (0.415 M-1 s-1) is only marginally lower than for 2-SMe (0.634 M-1 s-1) indicates that other factors must be important. One such factor may be π-donation by the MeS group in the substrate (12 and 13). According to

the pKCH values (Table 2) the electron-withdrawing effect a of the C(COO)2C(CH3)2 moiety is stronger than that of the O2NCCO2Me moiety, which probably translates into a stronger π-donor effect in the Meldrum’s acid derivative. Since π-donation stabilizes the substrate this will lead to a larger reduction in reactivity of 2-SMe than of 4-SMe. (4) In contrast to the k1(OH) values, the k1(pip) and k1(mor) values for 4-SMe are substantially lower than for 2-SMe: 39fold for k1(pip) and 29-fold for k1(mor), respectively. A possible explanation of this is that the transition state of the amine reactions is stabilized by intramolecular hydrogen bonding between the developing ammonium ion and the developing negative charge, which is akin to the intramolecular hydrogen bond in the fully developed T( A . The results suggest that this hydrogen bond and thus the transition state stabilization is stronger in the reaction of 2-SMe than in the reaction of 4-SMe. Such stronger hydrogen bonding could arise if charge development on the nitrogen were more advanced at the transition state of the reaction of 2-SMe. The larger βnuc value (0.41) for the reaction of 2-SMe compared to that of 4-SMe (0.33) is consistent with this notion. A larger steric effect in the reaction of 4-SMe may also contribute to the difference in the rates; even though such a steric effect would probably not discriminate between the k1(OH-) values due to the small size of the nucleophile, such discrimination is more plausible for k1(pip) and k1(mor) because of the bulkiness of the amines. PT (5) On the basis of the pKCH a and log k ° values, one would have expected 1-OMe to be much less reactive than 4-SMe or 2-SMe, but this is not the case. The main factor responsible for the higher than expected reactivity of 1-OMe is the replacement of the MeS group with the MeO group. In reactions with bulky nucleophiles such as secondary amines the smaller MeO group leads to a strong reduction in steric crowding at the transition state.2 This appears to be the main reason why k1(pip) and k1(mor) for 1-OMe are higher than the respective rate constants for 4-SMe. The enhanced electron-withdrawing inductive effect of the MeO group over the MeS group24 should also contribute to the higher rates for 1-OMe, although the stronger π-donor effect of the MeO group26 would probably offset this factor. For the reaction with OH-, steric effects should play a much smaller role. However, here the anomeric effect27 is responsible for the relatively high k1(OH) value for 1-OMe relative to that for 4-SMe and 2-SMe. Why Is the Intermediate Detectable in the Reactions with Piperidine and Morpholine? As stated in the Introduction, there is no accumulation of any intermediates to detectable levels (25) Hansch, C.; Leo, A.; Taft, R. W. Chem. ReV. 1991, 91, 165. (26) σR(OMe) ) -0.43; σR(SMe) ) -0.15.25 (27) In the present context, the anomeric effect28 refers to the stabilization exerted by geminal oxygen atoms,29 e.g., in dialkoxy or alkoxyhydroxy adducts.

3308 J. Org. Chem., Vol. 72, No. 9, 2007

in the reactions of 1-OMe or 2-SMe with strongly basic amines because only the thermodynamic condition (eq 3) but not the kinetic condition (eq 4) is met. However, in the reaction of 4-SMe with piperidine, T( A and TA do accumulate to detectable levels and so does TA in the reaction of 4-SMe with morpholine at high pH. Furthermore, in the reaction of 1-OMe with the much less basic amines methoxyamine and N-methyl9 methoxyamine, TA could also be observed directly. These contrasting observations are related to an important difference between the reactions of 4-SMe and 1-OMe with amines: In the reaction of 1-OMe the dominant pathway for the conversion of TA to products is the water-catalyzed leaving group departure (kH3 2O), whereas in the reaction of 4-SMe it is the acid-catalyzed pathway (kAH 3 ) that is dominant. Hence in the reaction of 1-OMe the strong electronic push (βpush ) 0.71) becomes a determining factor regarding the detectability of an intermediate. Specifically, the strong push leads to kH3 2O(pip) . kH3 2O(mor) . kH3 2O (MeONHMe) and, coupled with the small βnuc (0.25), to (k1/kH3 2O)(pip) < (k1/kH3 2O)(mor) < (k1/kH3 2O)(MeONHMe). This explains why intermediates derived from less basic amines are more easily detectable than those derived from more basic amines. In the reactions of 4-SMe the electronic push apparently plays a minor role if any; this can be seen from the fact that -1 s-1 is considerably larger than kAH(pip) kAH 3 (mor) ) 2.08 M 3 -2 -1 ) 4.25 × 10 M s-1 and is reflected in R ) 0.73 for acid catalysis.30 The larger kAH 3 (mor) value coupled with k1(pip) > k1(mor) yields (k1/kAH )(pip) . (k1/kAH 3 3 )(mor), i.e., here increased basicity of the amine enhances the detectability of the intermediate. This explains why TA accumulates to detectable levels with strongly basic amines and that TA derived from piperidine is more easily detectable than TA derived from morpholine. Regarding the reaction of 2-SMe with piperidine and morpholine, not enough information is available to pinpoint the reasons why the respective intermediates were not observable. For one, under conditions where nucleophilic attack is not ratelimiting, ammonium ion catalyzed leaving group departure is rate-limiting only for the piperidine reaction, whereas for the morpholine reaction deprotonation of T( A was found to be ratelimiting.8a Why Is the Intermediate in the Reaction with Hydroxide Ion Not Detectable? There is no evidence that the presumed intermediate in the reaction of 4-SMe with OH-, 14, ac-

cumulates to detectable levels. This finding is consistent with observations made earlier in the hydrolysis of 1-OMe,31 (28) (a) Kirby, A. G. The Anomeric Effect and Related Stereoelectronic Effects of Oxygen; Springer-Verlag: Berlin, 1983. (b) Schleyer, P. v. R.; Jemmis, E. D.; Spitznagel, G. W. J. Am. Chem. Soc. 1985, 107, 6393. (29) (a) Hine, J.; Klueppl, A. W. J. Am. Chem. Soc. 1974, 96, 2924. (b) Wiberg, K. B.; Squires, R. R. J. Chem. Thermodyn. 1979, 11, 773. (c) Harcourt, M. P.; More O’Ferrall, R. A. Bull. Soc. Chim. Fr. 1988, 407. (30) The βpush value cannot possibly be larger than 1 - R ) 1 - 0.73 ) 0.27; a βpush value of 0.27 would imply a “true” R value for acid catalysis of 1.0, which is unlikely. (31) Bernasconi, C. F.; Fassberg, J.; Killion, R. B., Jr.; Schuck, D. F.; Rappoport, Z. J. Am. Chem. Soc. 1991, 113, 4937.

SNV Reaction with Amine Nucleophiles

2-OMe,32 2-SMe,33 3-OMe2, and other highly activated substrates. As with these earlier examples, it is the kinetic rather than the thermodynamic condition that is not met. As observed before,31,33 this is because conversion of 14 to products is unusually fast due to the availability of additional pathways. One such pathway involves rapid deprotonation of the OH group, generating the dianionic form of two intermediate, 15, which expels the leaving group much more rapidly than the monoanionic form because of the extra push. The other pathway involves intramolecular acid catalysis of leaving group departure by the OH group.

k1(OH) value for 1-OMe is also higher than expected. Here the main factor appears to be the anomeric effect; because of the small size of the nucleophile, the smaller steric effect is probably of minor importance. Experimental Section The synthesis of 4-SMe involved the conversion of methyl phenylpropiolate (16) to methyl β-iodo-R-nitrocinnamate (4-I) (eq 16) followed by substitution of the iodo group by the MeS group (eq 17).

Conclusions (1) The reaction of 4-SMe with piperidine and morpholine is the first example of an SNV reaction with moderately to strongly basic amine nucleophiles where the intermediate accumulates to detectable levels. The only other SNV reaction involving amine nucleophiles that has allowed a direct observation of the intermediate is that of 1-OMe with methoxyamine and N-methylmethoxyamine, but these are weakly basic amines; in the reaction of the same substrate with morpholine or piperidine TA is a nonobservable steady state intermediate. (2) The reason why for 4-SMe it is easier to detect the intermediate in its reactions with highly basic amines whereas for 1-OMe TA is more easily observed in reactions with weakly basic amines is the fact that leaving group departure in the former reactions is catalyzed by the respective protonated amine but not in the latter. This leads to (k1/kAH 3 )(pip) . (k1/ )(mor) in the reactions of 4-SMe, but because βpush > βnuc, kAH 3 to (k1/kH3 2O)(pip) < (k1/kH3 2O)(mor) < (k1/kH3 2O)(MeONHMe) in the reactions of 1-OMe. (3) No intermediate could be observed in the reaction of 4-SMe with OH-. This is because the acidic properties of the hydroxyl group of the intermediate provide additional pathways for the loss of the leaving group. ( (4) The pK( a values of TA from the reaction of 4-SMe are ( consistent with the pKa of T( A from the reaction of 2-SMe with piperidine and considerably higher than the pK( a values of the respective amine adducts of 1-OMe. This is because of strong intramolecular bonding in the former and weak intramolecular hydrogen bonding in the latter; the reduced electron-withdrawing effect of the MeS group compared to that of the MeO group also contributes to the difference. PT (5) The expectation, based on the pKCH a and log k ° values, that k1(OH) for 4-SMe should be significantly lower than for 2-SMe is not met. A possible reason is weaker π-donation by the MeS group in 4-SMe due to the stronger electronwithdrawing effect of the (COO)2C(CH3)2 group. On the other hand, the k1(pip) and k1(mor) values for 4-SMe are substantially lower than for 2-SMe. Stronger transition state stabilization by intramolecular hydrogen bonding in the reactions of 2-SMe is the likely cause for this finding. (6) The main reason why the k1(pip) and k1(mor) for the reactions of 1-OMe are higher than the respective rate constants for 4-SMe and higher than expected based on the pKCH and a PT log k ° values is the reduced steric crowding at the transition state when the MeS group is replaced by a MeO group. The (32) Bernasconi, C. F.; Ketner, R. J.; Chen, X.; Rappoport, Z. J. Am. Chem. Soc. 1998, 120, 7461. (33) Bernasconi, C. F.; Schuck, D. F.; Ketner, R. J.; Weiss, M.; Rappoport, Z. J. Am. Chem. Soc. 1994, 116, 11764.

Synthesis of 4-I. Dinitrogen tetraoxide (2.6 mL, 0.04 mmol) was transferred by a stream of dry argon to a solution containing methyl phenylpropiolate [(8 g, 0.05 mol; 1H NMR (CDCl3) δ: 3.84 (3H, s, OMe), 7.36-7.39 (2H, m, Ph), 7.43-7.45 (1H, m, Ph), 7.577.59 (2H, m, Ph)] and iodine (20 g, 0.08 mol) in dry ether (300 mL) during 5 h at room temperature. The solution was then stirred for 40 h at room temperature. To the dark red solution was slowly added a 5% aqueous Na2S2O3 solution (6 × 250 mL) until the iodine color had completely disappeared. The organic phase was washed with water (2 × 100 mL) and dried (MgSO4). Evaporation of the solvent left a mixture of a solid and an oil (15 g, 90%). After washing with cold petroleum ether, a yellow solid (10 g), mp 98-104 °C, was obtained. An additional amount was obtained from the washing solution. The product was separated into the E and Z isomers by crystallization from petroleum ether or CHCl3 and gave 99% Z isomer, 1% E isomer in the best case. It was obtained as the pure isomer from chromatography over silica column using 90% petroleum ether, 40-60 °C, 10% EtOAc eluent, followed by crystallization from petroleum ether. Two consecutive crystallizations from petroleum ether gave pure E-isomer. Data for E-isomer: yellow crystals, mp 119-120 °C. 1H NMR (CDCl3) δ: 3.94 (3H, s, OMe), 7.28-7.45 (5H, m, Ar); (DMSO-d6) δ: 3.89 (3H, s, OMe), 7.29-7.47 (5H, m, Ar). MS: m/z (relative abundance %, assignment) 333 (8, M), 305 (3, M - CO), 176 (3, M - I NO), 160 (29, M - I - NO2), 129 (100, M - PhI), 105 (45, PhCO), 103 (4, PhCO - 2H), 102 (31, PhCO - 3H), 77 (14, Ph). Anal. Calcd for C10H8O4IN: C, 36.06; H, 2.42; N, 4.21. Found: C, 35.94; H, 2.38; N, 4.12. Data for Z-isomer: yellow crystals, mp 6565.5 °C. 1H NMR (CDCl3) δ: 3.63 (3H, s, MeO), 7.30-7.42 (5H, m, Ph); (DMSO-d6) δ: 3.56 (3H, s, OMe), 7.29-7.50 (5H, m, Ar). MS: m/z (relative abundance %, assignment) 333 (6, M), 305 (2, M - COO, 228 (s), 176 (3, M - I - NO), 160 (19, M - I NO2), 129 (100, M - Ph - I), 105 (63, PhCOO, 103 (5, PhCO 2H), 77 (25, Ph). Anal. Calcd. for C10H8O4IN: C, 36.06; H, 2.42; N, 4.21. Found: C, 36.27; H, 2.38, N, 4.27. Both pure isomers decompose on standing in DMSO-d6. Synthesis of 4-SMe. To a solution of a 5:1 E/Z mixture of 4-I (999 mg, 3 mmol) in acetonitrile (150 mL) was added the sparingly soluble sodium methylthiolate (252 mg, 3.6 mM). The turbid mixture was stirred for 4 h at room temperature until a precipitate was formed. Water (100 mL) was added, most of the MeCN was evaporated, and the residue was extracted with CHCl3 (3 × 100 mL). The CHCl3 solution was dried (MgSO4) and filtered, and evaporation of the solvent left a crude yellow oil (600 mg, 79%), higher field 1H NMR (CDCl3) δ 1.84 (E), 1.88 (Z) (2s, MeS, E/Z ratio ) 0.7, 29%), which contained a large percentage of the precursor methyl propiolate. The mixture was chromatographed over a silica column using 85% petroleum ether 40-60 °C/15% EtOAc as eluent. This ester was the main product eluted in the first fractions. J. Org. Chem, Vol. 72, No. 9, 2007 3309

Bernasconi et al. Middle fractions were mixtures of the propiolate ester and Eand Z-4-SMe. The last fractions were mixtures of E- and Z-4-SMe. Crystallization from petroleum ether gave a white solid (54 mg), mp 73-4 °C with E/Z ratio of 0.5. 1H NMR (CDCl3) δ: 1.84 (E), 1.88 (Z) [3H, 2s, MeS], 3.51 (Z), 3.88 (E) [3H, 2s, MeO], 7.207.48 [5H, m, Ar], E/Z ca. 0.5. MS: m/z (relative abundance %, assignment) 253 (7, M), 223 (5, M - NO), 221 (4, M - S), 174 (9, M - MeS - MeOH), 160 (23, M - MeS - NO2), 148 (14, M - CO2Me - NO2), 133 (M - CO2Me - NO2 - Me), 130 (10, M - Ph - NO2), 129 (M - Ph - MeS), 105 (100, PhCO), 102 (21, PhCO, 3H), 77 (38, Ph). Anal. Calcd for C11H11NO4S: C, 52.16; H, 4.38; N, 5.53. Found: C, 52.06, H, 4.33, N, 5.50. Additional fractions gave 13 mg of E/Z of ca. 0.7 and 28 mg of E/Z ca. 1.25, mp 90-91 °C. Altogether, 100 mg of E- and Z-4SMe (13.5%) was obtained. No attempt to optimize the yield was made. Attempts to separate further the E- and Z-isomers by crystallization from petroleum ether, CHCl3, and CH2Cl2 or by chromatography on silica gave in each experiment a different E/Z composition and separation was not achieved. The ratios were temperature- and solvent-dependent, e.g., E/Z ) 1.1 (CDCl3), 0.7

3310 J. Org. Chem., Vol. 72, No. 9, 2007

(DMSO-d6). A sample with E/Z ) 0.7 in DMSO-d6 at room temperature gave an E/Z ratio of 0.5 after 26 h at room temperature. Other Materials. Piperidine, morpholine, and triethylamine were distilled over sodium metal in an argon atmosphere. Solutions of 2 M KOH and 2 M HCl were prepared using Baker Dilut-it cartridges. DMSO was distilled over CaH2. Methodology. Preparation of solutions, pH measurements, recording of spectra, kinetic measurements, and data analysis were performed using general methods described previously.33

Acknowledgment. This research was supported by Grants CHE-9734822 and CHE-0446622 from the National Science Foundation (CFB), and a grant from the U.S.-Israel Binational Science Foundation (ZR). Supporting Information Available: Figures S1-S4 (kinetic data) and Tables S1-S2 (kinetic parameters). This material is available free of charge via the Internet at http://pubs.acs.org. JO062602S